Advances in Biochemistry & Applications in Medicine Chapter 12 Sirtuins: Its Role in Metabolic Homeostasis Midathala Raghavendra1; Naresh Kumar2; Jayanti Tokas1*; Hari Ram Singal2 1Department of Biochemistry, Advanced Post Graduate Centre (ANGRAU), Lam, Guntur, Andhra Pradesh, India 2Department of Chemistry and Biochemistry, CCS Haryana Agricultural University, Hisar, Haryana, India *Correspondence to: Jayanti Tokas, Department of Chemistry and Biochemistry, CCS Haryana Agricultural University, Hisar, Haryana, India Mob: +91-9729989988, Email : [email protected] Key Words: Isonicotinamide; Brown adipose tissue; Circadian clock; Hepatic glucose; Mitochondrial bio- genesis 1. Sirtuins: Its Role in Metabolic Homeostasis

Sirtuin are evolutionarily conserved enzymes that function in critical cellular processes such as DNA repair, transcription and stress resistance. The name Sirtuin (Sir2) is derived from the yeast ‘silent information regulator’ which is responsible for controlled expression of the silent mating type loci and also required for telomere hyper cluster formation in quiescent yeast cells [1]. It was not until fifteen years later the significance of sir2 proteins was identified as beta-nicotinamide adenine dinucleotide (NAD)-dependent deacety- lases, which deacetylate lysine at a specific site and accounts for silencing, recombination sup- pression, apoptosis, mitochondrial biogenesis, lipid metabolism, and extension of life span in- vivo [2,3]. The sirtuins were originally classified as histone deacetylases but many non-histone targets are described recently.

1.1. Biochemistry of sirtuins

Sirtuins possess a conserved catalytic core (~275 amino acids) that is flanked by N- and C- terminal extensions. These N- and C- terminal extensions play a key role in ensuring proper cellular localization, regulating the interaction with other proteins and targets for post-transla- tional modifications that affect the functions of sirtuins [4]. The larger sirtuin domain consists of a Rossmann-fold, that is characteristic for NAD+-binding unit, and the smaller domain: a Advances in Biochemistry & Applications in Medicine zinc-binding motif and a α-helical region which shows the highest diversity among family members [5]. Tokas J Tokas

Figure 1: structure of NAD-dependent protien Deacetylase Sirtuin-1 (PDB ID 41G9): A. Ball and stick model B. space fill model acetylation is regulated by protein acetyltransferases and deacetylases. Sirtuins are a family of NAD+ - dependent protein deacetylases (class III) which are widely distributed in almost all phyla of life. Sirtuins differ from other classes of deacetylases in that they are ab- solutely dependent on NAD+, deacetylates lysine residue and releases nicotinamide (NAM), 2’ O-acetyl-adenosine diphosphate-ribose (AADPR) and a deacetylated substrate. As sirtuins are dependent upon the presence of NAD+, sirtuin activity is directly linked to the metabolic state of the cell [6]. In the first step of the reaction, ADP-ribose is covalently attached to the acetyl moiety of the substrate, accompanied by the release of free NAM. Hydrolysis of the acetyl- lysine bond then occurs, liberating the AADPR. NAM acts as an inhibitor of the reaction and thus provides negative feedback inhibition of the sirtuins in vivo [7]. The byproduct O-acetyl ADP-ribose (AADPR) which is released in the deacetylation reaction has an essential role in modification of proteins [8,9]. Macrodomain, a conserved globular protein domain of 130-190 amino acids recognizes the terminal ADP-ribose of poly ADP ribose (PAR). Only 11 human macrodomain-containing proteins have been identified so far which are capable of binding to AADPR. Macrodomains are involved in DNA repair, transcriptional regulation, signalling events, control of NAD+ metabolism, chromatin remodelling and developmental processes [9]. www.openaccessebooks.com

Figure 2: the sir2 reaction. Deacetylation of protien acetyllysine catalysed by sir2p. Acetyl-group transfer to the ADP- ribose (ADPR) motely of NAD* occurs via sir2p chemistry to form 2’-o-acetly-ADPR. (Adopted from sauve et al. 2006 2 Advances in Biochemistry & Applications in Medicine Sirtuins are absolutely dependent on NAD+ the abundance of free NAD+ and its biosyn- thetic and breakdown products in cells are relevant to the enzymatic activity of sirtuins [10]. The root substrates for different NAD+ biosynthetic pathways include the amino acid trypto- phan (Trp), nicotinic acid, NAM and nicotinamide riboside (NR). The intermediate nicotin- amide mononucleotide (NMN) can also directly stimulate NAD+ synthesis [11]. Nicotinamide phosphoribosyltransferase (Nampt) catalysis is a rate-limiting step which transfer of phos- phoribosyl group from 5-phosphoribosyl-1-pyrophosphate to nicotinamide forming NMN and pyrophosphate. The cellular levels of Nampt which synthesize NAD+ vary during different pathophysiological conditions in metabolism like inflammation and cancer [12]. Increased expression of Nampt in response to various stresses elevates cellular NAD+ levels, which in turn regulate catalytic activity. For example, intracellular NAD levels regulate tumour necrosis factor protein synthesis in a sirtuin-dependent manner [13].

1.2. Modulators of sirtuins

The sirtuins promote longevity in diverse species and mediate many of the beneficial effects of calorie restriction (CR), such as reduced incidence of cancer, cardiovascular disease and diabetes. Sirtuins attracted considerable interest as a therapeutic target for the develop- ment of drug targets [7]. Many inhibitors and SIRT1-activating compounds (STACs) have been discovered for sirtuins but at the molecular level, the mechanism by which sirtuins is activated remains elusive [14].

Nicotinamide which is a byproduct of the deacetylation process by sirtuins acts as an inhibitor. The synthetic molecule iso-nicotinamide (iNAM) can act as a pan-sirtuin activa- tor within the limits of pharmacological concentrations. Several classes of plant-derived me- tabolites such as flavones, stilbenes, chalcones and anthocyanides activate SIRT1 in vitro. Resveratrol (3,5,4’ – trihydroxystilbene), is the most potent of the natural activators of the sirtuins. Unfortunately, resveratrol as a drug is not likely to succeed due to insolubility, it's poor bioavailability and rapid half-life [15]. The first synthetic STACs were derivatives of an imidazothiazole scaffold (e.g. SRT460, SRT1720 and SRT2183) and chemically distinct from the polyphenol backbone of resveratrol. More potent second-generation STACs based on ben- zimidazole and urea-based scaffolds were subsequently described [7,16].

The sirtuin inhibitors include sirtinol (reduces inflammation in capillary endothelial cells of the skin), cambinol (competitive inhibitor that competes with acetylated polypeptides), suramin (urea derivative competes for binding with both NAD+ and the acetylated lysine of the substrate), EX-527 (SIRT1 inhibitor, increase the acetylation of protein at K382 following the induction of DNA damage in human mammary epithelial cell and some tumor cell lines), oxyindole (SIRT2 inhibitor, inhibits α-tubulin deacetylation in mammary cell lines [5].

3 Advances in Biochemistry & Applications in Medicine

Figure 3: Activation mechanisam of sirtuins by isonicotinamide (INAM). In the presence of nicotinamide the deacetyla- tion reaction is slow because of depletion of the imidiate by nicotinamide reactivity. isonicotinamidecan bind the nico- tinamide pocket and prevent reaction reversal. The INAM-bound complex can complete deacetylation more effeciently, leading to activation of sirtuin catalytic activity in cell. (Adapted from “The Biochemistry of sirtuins” by sauve et al. 2006) 1.3. Localization of mammalian sirtuins

The seven mammalian sirtuins are nuclear encoded and ubiquitously expressed in hu- man tissues and occupy three different subcellular localization nucleus, cytoplasm and mito- chondria. SIRT 1, -2, -6, -7 are found in nucleus. SIRT1 and 2 are found in the cytoplasm as well. SIRT3, 4 and 5 are found in the mitochondria [5,17]. Sirt1 localization is predominantly nuclear but can be translocated to the cytoplasm in a cell-specific and cell-autonomous man- ner, in response to various physiological stimuli and disease states. The subcellular localiza- tion, enzymatic activity and the diverse functions of sirtuin isoforms are shown in Table-1. The various sirtuins have very diverse substrates which can be broken down into three major overlapping classes: transcriptional, apoptosis regulating and metabolic regulating. The sub- strates for sirtuins in Homo sapiens include (H1, H3, H4), P53, FOXO3a, MyoD PGC-1α and other transcription factors which are deacetylated with transcription activation. In most cases, these transcription factors control related to growth, cell cycle and apoptosis control [10]. Table 1: Mammalian isoforms of Sirtuins, its enzyme activity and function

Sirtuin Subcellular localization Enzymatic activity Function Reference SIRT1 Nucleus Cytoplasm Deacetylase • Formation of facultative and constitutive [18–26] chromatin • Mitochondrial biogenesis • Fatty acid oxidation • Repress adipocyte differentiation • Regulation of cholesterol and bile acid homeostasis • Stimulate gluconeogenesis • Inhibit apoptosis

4 Advances in Biochemistry & Applications in Medicine

SIRT2 Nucleus Cytoplasm Deacetylase • Cell cycle regulation through mitosis [27–33] Demyristoylase • Promotion of lipolysis in adipocytes • Inhibit adipocyte differentiation • Tumour suppression/promotion • Neurodegeneration SIRT3 Mitochondria Deacetylase • Regulation of mitochondrial activity [34–43] Decrotonylase • Protection against oxidative stress ADP-ribosyltrans- • Enhance adaptive thermogenesis ferase • Tumor suppression • Accelerate acetyl CoA conversion • Facilitate TCA cycle and mitochondrial energy production • Enhance fatty acid oxidation SIRT4 Mitochondria ADP-ribosyltrans- • Glucose metabolism [44–48] ferase • Aminoacid catabolism Lipoamidase • Tumor suppression • Repress aminoacid stimulated insulin SIRT5 Mitochondria Deacetylase • Enhance urea cycle [49–52] Demalonylase • Fatty acid metabolism Desuccinylase • Amino acid metabolism Deglutarylase SIRT6 Nucleus Deacetylase • Genomic stability / DNA repair and pre- [53–57] Deacylase vent against ageing related disorders ADP-ribosyltrans- • Promote apoptosis ferase • Reduce glycolysis and increase mitochon- drial respiration • Reduce inflammatory response SIRT7 Nucleus Deacetylase • Ribosome biogenesis [58–61] • Tumor promotion

2. Sirtuins Mediate Effects of Calorie Restriction

There are many studies showing that sirtuins mediate the effects of calorie restriction (CR) in mammals. Restricting calorie intake, a reduction of calories by 20-40% is known as calorie restriction which can increase lifespan in lower organisms [62]. The two hallmark of CR is metabolic reprogramming (towards oxidative metabolism so as to gain most energy during the restricted diet) and resistance to stress (particularly oxidative stress) [63]. Calorie restriction induces the expression levels of SIRT1 and SIRT5 similarly, loss of function muta- tion of specific sirtuins ablates specific outputs of CR. However, high-fat diet leads to the loss of SIRT1 in mice [64]. The transgenic overexpression of SIRT1 or STACs mitigates disease syndrome like CR, these include diabetes, neurodegenerative diseases, liver steatosis, bone loss and inflammation [63]. Different groups of neurons in the hypothalamus control mamma- lian physiology & energy homeostasis, including feeding behaviour, nutrients inputs, energy expenditure, physical activity, body temperature and central circadian control.

Sirt1 is absolutely dependent on NAD+ and function as nutrient/redox sensor, and its expression is regulated by changes in nutritional status and is involved in a wide range of

5 Advances in Biochemistry & Applications in Medicine metabolic processes. Nutrient sensors have the ability to sense and respond to fluctuations in environmental nutrient levels which characterize a fundamental requisite for life [65]. There are diverse nutrient sensor pathways detecting intracellular and extracellular levels of sugars, amino acids, lipids and other metabolites that incorporate and correspond at the organismal level through hormonal signals. During the period of food-richness, nutrient-sensing pathways engage in anabolism and energy storage. Conversely, shortage of food triggers homeostatic mechanisms including mobilization of internal stores through autophagy. The expression lev- els of SIRT1 increases upon CR in rodent and human tissues. Levels of NAD rise under CR- like conditions, which in turn induces expression of SIRT1 [66]. Thus, sirtuins act as a meta- bolic sensor by detecting fluctuations in the NAD+/NADH ratio in cells. When the amount of nutrients decrease, especially in the case of glucose, levels of NAD+ increase and sirtuins are then elevated [67]. Calorie restriction increases SIRT1 deacetylase activity in skeletal muscle in parallel with enhanced insulin-stimulated phosphoinositide 3 kinases (PI3K) signalling and glucose uptake.

3. Sirt1 in Control of Metabolic Homeostasis

The brain plays a critical role in the regulation of systemic energy homeostasis. The hypothalamus is directly sensitive to nutritional changes [68]. Among several key brain areas involved in the regulation of energy balance, the hypothalamus is the primary structure that interprets adiposity and nutrient-related inputs [69]. Among the most prominent regulators within the hypothalamus, neurons in the circumventricular organ called the arculate nucleus (ARC) of the hypothalamus located in the mediobasal hypothalamus, anteriorly juxtaposing the median eminence (ME) is of critical importance for the regulation of energy balance. Leptin is a hormone released from adipose tissue binds to leptin receptors (LEPR) on agouti- related protein (AGRP)- producing neurons and pro-opiomelanocortin (POMC)- producing neurons in the ARC of the hypothalamus [70,71]. During fasting circulating leptin levels de- cline rapidly. The fall in leptin stimulates the expression of AGRP and neuropeptide Y (NPY) and suppresses POMC and cocaine- and amphetamine-regulated transcript (CART), thereby increasing food intake and decreasing energy expenditure [72,73]. Recent studies have shown that SIRT1 may play a role in central energy regulatory area. For instance, it has been shown that both CR and fasting enhance SIRT1 expression and activity in the hypothalamus [69].

At the molecular level, FoxO1 is a metabolic sensor that integrates both leptin and in- sulin signalling. FoxO1 is the master regulator of energy metabolism across species. FoxO1 is one of the four FoxO isoforms of transcriptional factors and is highly expressed in insulin- responsive tissues including pancreas, liver, skeletal muscle and adipose tissue [74,75]. In all these tissues FoxO1 orchestrates the transcriptional cascades regulating glucose metabolism. During fasting, FoxO1 promotes adaptation by inducing gluconeogenesis in the liver and a transition from carbohydrate oxidation to lipid oxidation in the muscle. In the pancreas, FoxO1 6 Advances in Biochemistry & Applications in Medicine inactivation is required for β-cell proliferation. Insulin suppresses FoxO1 activity through ac- tivation of PI3K / AKT signalling pathway. Activated AKT (also known as protein kinase B) phosphorylates FoxO1 at three highly conserved phosphorylation sites resulting in its nuclear exclusion and thus inhibition of transcription [75].

The hypothalamic mTOR (mammalian target of rapamycin) plays a role in cellular en- ergetics by inducing numerous anabolic protein processes and lipid synthesis and it signals suppression of food intake. AMP-activated protein kinase (AMPK) is a serine/threonine ki- nase which is also a nutrient /energy sensor whose activity is coupled to the energy status of the cells. AMPK is activated by increased AMP/ATP ratio that occurs during fasting [76]. AMPK is involved in the regulation of numerous biochemical pathways to turn off anabolism including fatty acid degradation. AMPK is an important regulator of energy homeostasis and is stimulated by a decrease in cellular energy status, nutrient and oxygen deprivation and in- creased energy expenditure. Activation of AMPK results in increased expression of Nampt and supply of NAD+ to support SIRT1 activity [77,78]. AMPK might phosphorylate SIRT1, disrupting the interaction with its negative regulator DBC1 (Deleted in Breast Cancer1, also known as KIAA1967).

Figure 4: Mechanism regulating FOXO transcription factors. Acetylation of FOXO transcription factor is retained in the nucelus, wheresas, phosphorylation excludes the FOXO from nucleus to cytoplasm. ubiquitin dependent degradation is an irreversible step. Ub, ubiquitin; p, phosphate group; Ac, acetyl group.

7 Advances in Biochemistry & Applications in Medicine 3.1. Lipid metabolism

In mammals, most energy is stored as fat in adipose tissues. White adipose tissue (WAT) is the main “storage site” of excess energy, primarily in the form of triglycerides. During fast- ing, WAT releases fatty acids that are used as fuel by other organs [21,79]. In addition, a func- tionally and morphologically distinct adipocyte with dense mitochondria was found in Brown adipose tissue (BAT). Brown adipose tissue dissipates energy as heat (non-shivering thermo- genesis). Brown adipocytes uncouple mitochondrial electron transport from ATP synthesis to a greater extent permeating the inner mitochondrial membrane to allow inter-membrane proton to leak back into the mitochondrial matrix, primarily through uncoupling protein-1 (UPC1) [80,81]. One of the critical regulators of fat storage in WAT is the nuclear peroxisome proliferating-activated receptor- γ (PPAR γ), whose activity promotes adipocyte differentia- tion and lipid anabolism. One of the suggested mechanism for loss of SIRT1 in obese animals is suggested by the finding that a high-fat diet in mice triggers cleavage of SIRT1 in WAT by caspase1 of the inflammasome [63]. The gain of function of NAD-dependent deacetylase SIRT1 or loss of function of its endogenous inhibitor DBC1 promotes “browning” of WAT by deacetylating PPAR- γ on K268 and K293 [82].

During fasting SIRT1 associated with PPAR γ and promoted the binding of the nuclear receptor corepressor1 (NCoR1). SIRT1 directly deacetylates PPAR γ which allows the recruit- ment of PRDM 16 to drive browning of white fat. Brown adipocytes are characterized by the expression of mitochondrial uncoupling protein 1 (UCP1), which allows dissipation of energy as heat for thermogenesis. The binding of PGC-1α to PPARγ promotes brown adipocyte-like features in white adipocytes through an upregulation of brown-adipocyte specific genes, such as UCP1, and a down-regulation of white-adipocyte specific genes. SIRT1 activation might prevent excessive accumulation of fat in adipocytes by boosting fat consumption and enhanc- ing thermogenic function.

SIRT1 inhibits lipogenic gene expression by acting as a negative regulator of the Sterol Regulatory Element Binding Protein (SREBP)-1c. SREBP-1c is a transcription factor that pro- motes the expression of lipogenic and cholesterogenic genes in order to facilitate fat storage [83]. The deacetylation of SREBP-1c by SIRT1 renders the protein susceptible to ubiquitin- mediated degradation [84]. Hence, SIRT1 activation leads to decreased SREBP-1c protein levels. This results in decreased occupancy of SREBP-1c on the promoter of lipogenic genes and a concomitant reduction in their expression levels. Deacetylation of SREBP1 by SIRT1 results in targeting of SREBP1 for proteasomal degradation, which inhibits the expression of lipogenic and cholesterol synthesis genes [85]. SIRT1 also promotes reverse cholesterol trans- port by deacetylating and activating the LXRα nuclear receptor, and promotes the cholesterol catabolic pathway by deacetylating and activating BAR, bile acid receptor; LXRα, oxysterols receptor; SREBP-1, sterol regulatory element binding protein 1 [86,87] 8 Advances in Biochemistry & Applications in Medicine Indeed, SIRT1 has been shown to modulate cholesterol metabolism in vivo through positive regulation of the Farnesoid X receptor (FXR) and the Liver X receptors (LRX), LXRα and LXRβ. In FXR, SIRT1 can directly deacetylate Lys 157 and Lys 217. Down-regulation of hepatic SIRT1 increases FXR acetylation, which inhibits its heterodimerization with the Retinoid X receptor (RXR) α and therefore, its transcriptional activity [88,89]. Hence, SIRT1 deletion in the liver is sufficient to downregulate FXR-related transcriptional programs and lead to the formation of cholesterol gallstones [23,86]. In LXR, ligand binding promotes the interaction with SIRT1 and subsequent deacetylation on Lys 432 (LXR α) and on Lys 433 (LXR β), promoting their activation.

Figure 5: the liver X receptors (LXRs) and retinoid X receptors (RXRs) from heterodimer and blind to direct repeat 4 (DR4) response element in regulatory regions of their target genes. A. In basal state co-repressors are bound to heterodi- mer, which inhibits transcription of target genes. B. when RXR and LXR are activated by biniding of retinoic acid and oxysterols respectively the co-activators are recruited and initiate transcription of target gene. Liver X receptor (LXR) forms an obligate heterodimer with retinoid X receptor (RXR) that binds to a DR4 (direct repeat spaced by four nucleotides) LXRE (LXR response element) in the regulatory regions of target genes, thereby repressing gene expression [86,90]. Follow- ing ligand binding to LXR or RXR, the heterodimer changes conformation, this leads to the release of corepressors and the recruitment of coactivators. This results in the transcription of target genes. Similarly, farnesoid X receptor (FXR) forms a heterodimer with RXR and binds to the FXR response element (FXRE), which is typically an inverse repeat spaced by one nu- cleotide (IR1), in its target genes to induce gene expression. Knockdown of SIRT1 in the liver also leads to decreased expression of CYP7A1, a bonafide LXR target. LXRs are also a potent inducer of lipid anabolism by increasing SREBP-1c activity. However, SIRT1 can deacetylate SREBP-1c resulting in proteasomal degradation [83]. The liver X receptor (LXR) controls both sterol regulatory element–binding protein (SREBP)-2 and SREBP-1c. SREBP-2 regu- lates the genes involved in cholesterol synthesis, while SREBP-1c stimulates the production of genes involved with the lipogenic enzymes. Inhibition of LXR would result in a decrease in both cholesterol and triglyceride synthesis [90].

Therefore, SIRT1 activation might promote the beneficial effects of LXR activity on cholesterol homeostasis while preventing the detrimental effects on lipid anabolism by deacetylating SREBP-1c. Altogether, SIRT1 overexpression improves cholesterol metabolism and prevents hepatic steatosis, while SIRT1 deletion in the liver favours lipid accumulation.

9 Advances in Biochemistry & Applications in Medicine 3.2. Hepatic glucose metabolism

SIRT1 participate in the energy regulation in metabolically active tissues such as liv- er and muscle. Gluconeogenesis during starvation can be triggered by the hormone gluca- gon, which induces dephosphorylation and nuclear translocation of the transducer of regu- lated CREB activity2 (TORC2) [91]. TORC2 activates DNA transcription factor Cre binding protein (CREB), which then induces the expression of the transcription coactivator PGC-1α. PGC-1α complexes with transcription factor PPARα, FOXO1, glucocorticoid receptor (GR) and hepatocyte nuclear factor 4α (HFN4 α) [92]. These, in turn, induce the transcription of key gluconeogenic genes encoding the rate-limiting enzymes such as phosphoenolpyruvate carboxykinase and glucose-6-phosphatase [93].

In the fed state, PCG1α is phosphorylated by insulin-stimulated Akt kinase activity thereby reducing the transcription of gluconeogenic genes. At low levels of cellular ATP, in liver, AMPK inhibits gluconeogenesis by phosphorylating TORC2. AMPK activation is a key pharmacological target for developing drugs for type-II diabetes which work to suppress he- patic glucose output such as metformin [92,94]. Besides phosphorylation, PGC1 α is also regulated by acetylation and deacetylation. PGC1α is activated by SIRT1-mediated deacety- lation during prolonged fasting which results in fatty acid oxidation and improved glucose homeostasis. This regulation of PGC1α activity is not dependent on glucagon or glucocorti- coids but on the levels of metabolic intermediates like pyruvate and NAD. SIRT1 also affect gluconeogenic activity through PGC1α in an indirect manner. In the liver, STAT3 is known to suppress the expression of PGC1α and gluconeogenic activity gene expression. Regulation of gluconeogenesis by STAT3 could be linked to nutritional status via SIRT1, which can directly deacetylate and attenuate the anti-gluconeogenic transcriptional activity of STAT3 [25,92,95]. SIRT1 also deacetylates and activates transcriptional factor Foxo1, resulting in increased glu- coneogenesis [96].

10 Advances in Biochemistry & Applications in Medicine Figure 6: SIRT1 mediates metabolic benefits in various tissues. Metabolically active tissues such as liver, heart, White adipose tissue (WAT) and skeletal muscle are shown along with SIRT1 specific function. PPAR-γ (peroxisome pro- liferator-activated receptor, PGC-1α (PPAR-γ coactivator 1α, FOXO1 (Forkhead box O1), CRTC2 (CREB-regulated transcription coactivator 2), PGAM-1 (Enzyme phosphoglycerate mutase-1), PPARα (peroxisome proliferator-activated receptor α), SREBP1c (sterol regulatory element binding protien 1c), AMPK (AMP-activated protein kinase), NAMPT (nicotinamide phosphorilbosylt transferase), LKB1 (Liver kinase B1). Prdm16 (PR domain containing 16), eNOS (en- dothelial nitric oxide synthase). Targets that are directly activated by SIRT1 are shown in Green. Those repressed or inhibited by SIRT1 are shown in pink. (Adopted from “SIRT1 and other sirtuins in metabolism” Chang et al. 2014) 4. Mitochondrial Biogenesis

SIRT1 has been described many times as a key regulator of mitochondrial biogenesis through the deacetylation of PGC-1α. PGC-1α becomes deacetylated in skeletal muscle dur- ing fasting. Similarly, PGC-1 α becomes deacetylated after exercise. Indeed, the activation of SIRT1 in response to nutrient or energy deprivation depends on AMPK activation, both in vitro and in vivo. AMPK is a master regulator of mitochondrial biogenesis and seems to play a key role in triggering SIRT1 activity during energy stress in skeletal muscle [97–99]. Cel- lular metabolic function controlled by mitochondrial number and function is modulated by SIRT1 through deacetylation of PPAR-co-activator 1 (PGC1). AMPK activation induces fatty acid oxidation in liver and heart, inhibits hepatic lipogenesis and adipocytes differentiation, and stimulates glucose uptake and mitochondrial biogenesis to modulate energy balance in muscle.

Insulin resistance is often characterised as the most critical factor contributing to the development of type 2 diabetes. Development of insulin resistance is a multi-step process with strong genetic and environmental influences, and the precise pathogenesis and pathophysi- ological sequence resulting in insulin resistance are still largely unknown. As a negative regu- lator of the insulin signal transduction cascade, PTP1B has been shown to function as a key in- sulin receptor phosphatase [100]. SIRT1 also deacetylates liver X receptor and downregulates protein tyrosine phosphatase 1B (PTP1b), a major tyrosine phosphatase for the insulin recep- tor and the insulin receptor substrate proteins, IRS1 and IRS2. SIRT1 was found to be required for the deacetylation and inactivation of the transcription factor STAT3 during CR, thereby promoting more efficient PI3K signalling during insulin stimulation [101]. SIRT1 could po- tentiate gluconeogenesis by directly deacetylating and enhancing the transcriptional activity of the Peroxisome Proliferator-activated Receptor (PPAR) gamma coactivator 1α (PGC-1α ) or the Forkhead O-box protein 1 (FoxO1) transcription factor, both considered key positive controllers of the gluconeogenic transcriptional program.

5. Regulation of Circadian Rhythm

Circadian rhythm refers to an endogenous entrainable 24-hour oscillation of any bio- logical process in all living organism. This circadian rhythm depends on internal clocks that

11 Advances in Biochemistry & Applications in Medicine work in part through chromatin modification and epigenetic control of gene expression [69]. The autonomous and self-sustainable nature of circadian timing is largely dependent on the molecular circadian clockwork. Cells modify their biochemical activities depending upon the food intake and energy expenditure. This can be achieved by fine tuning the central biochemi- cal pathway by various metabolic targets- mainly rate-limiting enzymes. The expression of these metabolic targets is modulated by the chromatin conformation and the accessibility of transcription factors that encode these enzymes. In turn, these chromatin conformations are controlled by histone acetylation, the levels of which are controlled by the concerted enzymat- ic activity of HATs and Histone deacetylases [102]. The circadian rhythms are directly dictated by the food availability, which is an external cue that entrains peripheral clocks. Further, sev- eral stimuli, including insulin, glucose, the glucocorticoid hormone analogue dexamethasone, forskolin and phorbol ester can trigger circadian expression in vitro by activating signalling cascades [103].

The molecular mechanism underlying the mammalian circadian clock consists of a tran- scriptional-translation feedback loop involving CLOCK (circadian locomotor output cycles ka- put) and BMAL1 (brain and muscle ARNT-like1), which recognize E-box elements. CLOCK, a transcription factor is crucial for circadian function, has intrinsic histone acetyl transferase (HAT) activity [102]. The CLOCK/BMAL1 heterodimer activates the transcription of the pe- riod (Per1, 2 and 3) and cryptochrome (Cry1 and 2) leading to the subsequent repression of CLOCK/BMAL1 activity. Indeed, the cellular DNA-binding activity of CLOCK/BMAL1 is strongly influenced by the ratio of reduced to oxidized NAD cofactors. SIRT1 is a component of CLOCK/BMAL1 transcription complexes and affects the expression of clock genes [104]. SIRT1 interacts with CLOCK-BMAL1 to control the amplitude and extent of circadian clock- controlled gene expression through deacetylation of PER2 and BMAL1. Circadian regulation of SIRT1 activity is due to circadian oscillations of the cellular NAD+ levels. It has been found that Nampt is a direct transcriptional target of CLOCK-BMAL1. Therefore, the feedback loop in the circadian clock that involves CLOCK-BMAL1, Nampt, NAD+ and SIRT1 provide an important connection between physiological rhythm and cellular metabolism.

Figure 7: CLOCK is associated in a nuclear complex with BMAL1 and SIRT1, an HDAC that is directly regulated by cellular metabolism. This comples is recruited to circadian gene promoters in a cyclic manner and is thought to be responsible for the circadian acetylation of histone H3 at K9 and K14. (Adopted from “Mammalian circadian clock and metabolism-the epigenetic link” Bellet & Sassone-Corsi 2010)

12 Advances in Biochemistry & Applications in Medicine 6. Summary

Sirtuins are a conserved protein with NAD+- dependent deacetylase activity and their functions are intrinsically linked to cellular metabolism. Sirtuins in lower organisms prolong the life and regulate the ageing process against metabolic stresses. The versatile functions of seven sirtuins are sustained by the distribution within the cellular compartment and tissues allowing cells to sense nutrient levels. SIRT1 plays a critical role in maintaining metabolic ho- meostasis with systemic effects via the hypothalamus and helps to deliver the benefits of calo- rie restriction. Sirtuins have numerous targets in many tissues such as liver, muscle, adipose tissue etc. which perceive the nutrient levels and respond through deacetylation of histones, key transcription factors and metabolic enzymes. SIRT1 networks with CLOCK-BMAL1 to regulate the physiological rhythm and cellular metabolism.

7. References

1. Rine J, Herskowitz I. Four genes responsible for a position effect on expression from HML and HMR in Saccharo- myces cerevisiae. Genetics. 1987; 116: 9–22.

2. Imai S-I, Armstrong CM, Kaeberlein M, et al. Transcriptional silencing and longevity protein Sir2 is an NAD-depen- dent histone deacetylase. Nature. 2000; 403: 795.

3. Poulose N, Raju R. Sirtuin regulation in aging and injury. Biochim Biophys Acta BBA-Mol Basis Dis. 2015; 1852: 2442–2455.

4. Moniot S, Weyand M, Steegborn C. Structures, substrates, and regulators of mammalian Sirtuins–opportunities and challenges for drug development. Front Pharmacol; 3.

5. Kupis W, Pałyga J, Tomal E, et al. The role of sirtuins in cellular homeostasis. J Physiol Biochem. 2016; 72: 371– 380.

6. Sebastián C, Satterstrom FK, Haigis MC, et al. From sirtuin biology to human diseases: an update. J Biol Chem. 2012; 287: 42444–42452.

7. Hubbard BP, Sinclair DA. Small molecule SIRT1 activators for the treatment of aging and age-related diseases. Trends Pharmacol Sci. 2014; 35: 146–154.

8. Leung AK, Todorova T, Ando Y, et al. Poly (ADP-ribose) regulates post-transcriptional gene regulation in the cyto- plasm. RNA Biol. 2012; 9: 542–548.

9. Hottiger MO. Nuclear ADP-ribosylation and its role in chromatin plasticity, cell differentiation, and epigenetics. Annu Rev Biochem. 2015; 84: 227–263.

10. Sauve AA, Wolberger C, Schramm VL, et al. The biochemistry of sirtuins. Annu Rev Biochem. 2006; 75: 435– 465.

11. Canto C, Menzies KJ, Auwerx J. NAD+ metabolism and the control of energy homeostasis: A balancing act between mitochondria and the nucleus. Cell Metab. 2015; 22: 31–53.

12. Garten A, Petzold S, Körner A, et al. Nampt: linking NAD biology, metabolism and cancer. Trends Endocrinol Me- tab. 2009; 20: 130–138.

13. Van Gool F, Galli M, Gueydan C, et al. Intracellular NAD levels regulate tumor necrosis factor protein synthesis in

13 Advances in Biochemistry & Applications in Medicine a sirtuin-dependent manner. Nat Med. 2009; 15: 206–210.

14. Dai H, Kustigian L, Carney D, et al. SIRT1 activation by small molecules kinetic and biophysical evidence for direct interaction of enzyme and activator. J Biol Chem 2010; 285: 32695–32703.

15. Massudi H, Wu LE, Sinclair DA. Sirtuin Activation by Small Molecules. In: Sirtuins. Springer, 2016, pp. 243–266.

16. Hubbard BP, Gomes AP, Dai H, et al. Evidence for a common mechanism of SIRT1 regulation by allosteric activa- tors. Science. 2013; 339: 1216–1219.

17. Kelly G. A review of the sirtuin system, its clinical implications, and the potential role of dietary activators like res- veratrol: part 1. Altern Med Rev. 2010; 15: 245–263.

18. Vaquero A, Scher M, Erdjument-Bromage H, et al. SIRT1 regulates the histone methyl-transferase SUV39H1 during heterochromatin formation. Nature. 2007; 450: 440–444.

19. Aquilano K, Vigilanza P, Baldelli S, et al. Peroxisome Proliferator-activated Receptor γ Co-activator 1α (PGC-1α) and (SIRT1) Reside in Mitochondria POSSIBLE DIRECT FUNCTION IN MITOCHONDRIAL BIOGEN- ESIS. J Biol Chem. 2010; 285: 21590–21599.

20. Gerhart-Hines Z, Rodgers JT, Bare O, et al. Metabolic control of muscle mitochondrial function and fatty acid oxida- tion through SIRT1/PGC-1α. EMBO J. 2007; 26: 1913–1923.

21. Picard F, Kurtev M, Chung N, et al. Sirt1 promotes fat mobilization in white adipocytes by repressing PPAR-γ. Na- ture. 2004; 429: 771.

22. Wang H, Qiang L, Farmer SR. Identification of a domain within peroxisome proliferator-activated receptor γ regu- lating expression of a group of genes containing fibroblast growth factor 21 that are selectively repressed by SIRT1 in adipocytes. Mol Cell Biol. 2008; 28: 188–200.

23. Kemper JK. Regulation of FXR transcriptional activity in health and disease: Emerging roles of FXR cofactors and post-translational modifications. Biochim Biophys Acta BBA-Mol Basis Dis. 2011; 1812: 842–850.

24. Purushotham A, Schug TT, Xu Q, et al. Hepatocyte-specific deletion of SIRT1 alters fatty acid metabolism and re- sults in hepatic steatosis and inflammation. Cell Metab. 2009; 9: 327–338.

25. Nie Y, Erion DM, Yuan Z, et al. STAT3 inhibition of gluconeogenesis is downregulated by SirT1. Nat Cell Biol. 2009; 11: 492.

26. Caton PW, Nayuni NK, Kieswich J, et al. Metformin suppresses hepatic gluconeogenesis through induction of SIRT1 and GCN5. J Endocrinol. 2010; 205: 97–106.

27. Serrano L, Martínez-Redondo P, Marazuela-Duque A, et al. The tumor suppressor SirT2 regulates cell cycle pro- gression and genome stability by modulating the mitotic deposition of H4K20 methylation. Genes Dev. 2013; 27: 639–653.

28. Peck B, Chen C-Y, Ho K-K, et al. SIRT inhibitors induce cell death and p53 acetylation through targeting both SIRT1 and SIRT2. Mol Cancer Ther. 2010; 9: 844–855.

29. Kim H-S, Vassilopoulos A, Wang R-H, et al. SIRT2 maintains genome integrity and suppresses tumorigenesis through regulating APC/C activity. Cancer Cell. 2011; 20: 487–499.

30. Ye X, Li M, Hou T, et al. Sirtuins in glucose and lipid metabolism. Oncotarget. 2017; 8: 1845.

31. Park S-H, Zhu Y, Ozden O, et al. SIRT2 is a tumor suppressor that connects aging, acetylome, cell cycle signaling, and carcinogenesis. Transl Cancer Res. 2012; 1: 15.

32. Luthi-Carter R, Taylor DM, Pallos J, et al. SIRT2 inhibition achieves neuroprotection by decreasing sterol biosyn- 14 Advances in Biochemistry & Applications in Medicine thesis. Proc Natl Acad Sci. 2010; 107: 7927–7932.

33. Donmez G, Outeiro TF. SIRT1 and SIRT2: emerging targets in neurodegeneration. EMBO Mol Med. 2013; 5: 344–352.

34. Hirschey MD, Shimazu T, Goetzman E, et al. SIRT3 regulates mitochondrial fatty-acid oxidation by reversible en- zyme deacetylation. Nature. 2010; 464: 121–125.

35. Verdin E, Hirschey MD, Finley LW, et al. Sirtuin regulation of mitochondria: energy production, apoptosis, and signaling. Trends Biochem Sci. 2010; 35: 669–675.

36. Bause AS, Haigis MC. SIRT3 regulation of mitochondrial oxidative stress. Exp Gerontol. 2013; 48: 634–639.

37. Bell EL, Guarente L. The SirT3 divining rod points to oxidative stress. Mol Cell. 2011; 42: 561–568.

38. Tseng AH, Shieh S-S, Wang DL. SIRT3 deacetylates FOXO3 to protect mitochondria against oxidative damage. Free Radic Biol Med. 2013; 63: 222–234.

39. Giralt A, Villarroya F. SIRT3, a pivotal actor in mitochondrial functions: metabolism, cell death and aging. Biochem. J 2012; 444: 1–10.

40. Shi T, Fan GQ, Xiao SD. SIRT3 reduces lipid accumulation via AMPK activation in human hepatic cells. J Dig Dis. 2010; 11: 55–62.

41. Schwer B, Bunkenborg J, Verdin RO, et al. Reversible lysine acetylation controls the activity of the mitochondrial enzyme acetyl-CoA synthetase 2. Proc Natl Acad Sci. 2006; 103: 10224–10229.

42. Hallows WC, Yu W, Smith BC, et al. Sirt3 promotes the urea cycle and fatty acid oxidation during dietary restriction. Mol Cell. 2011; 41: 139–149.

43. Bause AS, Haigis MC. SIRT3 regulation of mitochondrial oxidative stress. Exp Gerontol. 2013; 48: 634–639.

44. Jeong SM, Xiao C, Finley LW, et al. SIRT4 has tumor-suppressive activity and regulates the cellular metabolic re- sponse to DNA damage by inhibiting mitochondrial glutamine metabolism. Cancer Cell. 2013; 23: 450–463.

45. Csibi A, Fendt S-M, Li C, et al. The mTORC1 pathway stimulates glutamine metabolism and cell proliferation by repressing SIRT4. Cell. 2013; 153: 840–854.

46. Haigis MC, Mostoslavsky R, Haigis KM, et al. SIRT4 inhibits glutamate dehydrogenase and opposes the effects of calorie restriction in pancreatic β cells. Cell. 2006; 126: 941–954.

47. Miyo M, Yamamoto H, Konno M, et al. Tumour-suppressive function of SIRT4 in human colorectal cancer. Br J Cancer. 2015; 113: 492–499.

48. Argmann C, Auwerx J. Insulin secretion: SIRT4 gets in on the act. Cell. 2006; 126: 837–839.

49. Nakagawa T, Lomb DJ, Haigis MC, et al. SIRT5 Deacetylates carbamoyl phosphate synthetase 1 and regulates the urea cycle. Cell. 2009; 137: 560–570.

50. Nakagawa T, Guarente L. Urea cycle regulation by mitochondrial sirtuin, SIRT5. Aging. 2009; 1: 578.

51. Rardin MJ, He W, Nishida Y, et al. SIRT5 regulates the mitochondrial lysine succinylome and metabolic networks. Cell Metab. 2013; 18: 920–933.

52. Du J, Zhou Y, Su X, et al. Sirt5 is a NAD-dependent protein lysine demalonylase and desuccinylase. Science 2011; 334: 806–809.

53. Yang B, Zwaans BM, Eckersdorff M, et al. The sirtuin SIRT6 deacetylates H3 K56Ac in vivo to promote genomic

15 Advances in Biochemistry & Applications in Medicine stability. Cell Cycle 2009; 8: 2662–2663.

54. Mostoslavsky R, Chua KF, Lombard DB, et al. Genomic instability and aging-like phenotype in the absence of mam- malian SIRT6. Cell. 2006; 124: 315–329.

55. Mao Z, Hine C, Tian X, et al. SIRT6 promotes DNA repair under stress by activating PARP1. Science. 2011; 332: 1443–1446.

56. Zhong L, D’Urso A, Toiber D, et al. The histone deacetylase Sirt6 regulates glucose homeostasis via Hif1α. Cell. 2010; 140: 280–293.

57. Xiao C, Wang R-H, Lahusen TJ, et al. Progression of chronic liver inflammation and fibrosis driven by activation of c-JUN signaling in Sirt6 mutant mice. J Biol Chem. 2012; 287: 41903–41913.

58. Tsai Y-C, Greco TM, Cristea IM. Sirtuin 7 plays a role in ribosome biogenesis and protein synthesis. Mol Cell Pro- teomics. 2014; 13: 73–83.

59. Shin J, He M, Liu Y, et al. SIRT7 represses Myc activity to suppress ER stress and prevent fatty liver disease. Cell Rep. 2013; 5: 654–665.

60. Malik S, Villanova L, Tanaka S, et al. SIRT7 inactivation reverses metastatic phenotypes in epithelial and mesen- chymal tumors. Sci Rep; 5.

61. Yu H, Ye W, Wu J, et al. Overexpression of sirt7 exhibits oncogenic property and serves as a prognostic factor in colorectal cancer. Clin Cancer Res 2014; 20: 3434–3445.

62. Brooks CL, Gu W. How does SIRT1 affect metabolism, senescence and cancer? Nat Rev Cancer. 2009; 9: 123.

63. Guarente L. Calorie restriction and sirtuins revisited. Genes Dev. 2013; 27: 2072–2085.

64. Pfluger PT, Herranz D, Velasco-Miguel S, et al. Sirt1 protects against high-fat diet-induced metabolic damage. Proc Natl Acad Sci. 2008; 105: 9793–9798.

65. Efeyan A, Comb WC, Sabatini DM. Nutrient sensing mechanisms and pathways. Nature 2015; 517: 302.

66. Duan W. Sirtuins: from metabolic regulation to brain aging. Front Aging Neurosci; 5.

67. López-Lluch G. Mitochondrial activity and dynamics changes regarding metabolism in ageing and obesity. Mech Ageing Dev. 2017; 162: 108–121.

68. Nillni EA. The metabolic sensor Sirt1 and the hypothalamus: Interplay between peptide hormones and pro-hormone convertases. Mol Cell Endocrinol. 2016; 438: 77–88.

69. Li X. SIRT1 and energy metabolism. Acta Biochim Biophys Sin. 2013; 45: 51–60.

70. Varela L, Horvath TL. Leptin and insulin pathways in POMC and AgRP neurons that modulate energy balance and glucose homeostasis. EMBO Rep. 2012; 13: 1079–1086.

71. Bjorbæk C, Kahn BB. Leptin signaling in the central nervous system and the periphery. Recent Prog Horm Res. 2004; 59: 305–332.

72. Park H-K, Ahima RS. Physiology of leptin: energy homeostasis, neuroendocrine function and metabolism. Metabo- lism. 2015; 64: 24–34.

73. Yu C-H, Chu S-C, Chen P-N, et al. Participation of ghrelin signalling in the reciprocal regulation of hypothalamic NPY/POMC-mediated appetite control in amphetamine-treated rats. Appetite. 2017; 113: 30–40.

74. Gross D, Van Den Heuvel A, Birnbaum M. The role of FoxO in the regulation of metabolism. Oncogene. 2008; 27:

16 Advances in Biochemistry & Applications in Medicine 2320.

75. Kousteni S. FoxO1, the transcriptional chief of staff of energy metabolism. Bone. 2012; 50: 437–443.

76. Hardie DG. Sensing of energy and nutrients by AMP-activated protein kinase. Am J Clin Nutr. 2011; 93: 891S– 896S.

77. Koltai E, Szabo Z, Atalay M, et al. Exercise alters SIRT1, SIRT6, NAD and NAMPT levels in skeletal muscle of aged rats. Mech Ageing Dev. 2010; 131: 21–28.

78. Cantó C, Gerhart-Hines Z, Feige JN, et al. AMPK regulates energy expenditure by modulating NAD+ metabolism and SIRT1 activity. Nature. 2009; 458: 1056.

79. Chen D, Bruno J, Easlon E, et al. Tissue-specific regulation of SIRT1 by calorie restriction. Genes Dev. 2008; 22: 1753–1757.

80. Lomb DJ, Laurent G, Haigis MC. Sirtuins regulate key aspects of lipid metabolism. Biochim Biophys Acta BBA- Proteins Proteomics. 2010; 1804: 1652–1657.

81. Shi T, Wang F, Stieren E, et al. SIRT3, a mitochondrial sirtuin deacetylase, regulates mitochondrial function and thermogenesis in brown adipocytes. J Biol Chem 2005; 280: 13560–13567.

82. Qiang L, Wang L, Kon N, et al. Brown remodeling of white adipose tissue by SirT1-dependent deacetylation of Pparγ. Cell. 2012; 150: 620–632.

83. Ponugoti B, Kim D-H, Xiao Z, et al. SIRT1 deacetylates and inhibits SREBP-1C activity in regulation of hepatic lipid metabolism. J Biol Chem. 2010; 285: 33959–33970.

84. Boutant M, Cantó C. SIRT1 metabolic actions: integrating recent advances from mouse models. Mol Metab. 2014; 3: 5–18.

85. Schug TT, Li X. Sirtuin 1 in lipid metabolism and obesity. Ann Med 2011; 43: 198–211.

86. Li X, Zhang S, Blander G, et al. SIRT1 deacetylates and positively regulates the nuclear receptor LXR. Mol Cell. 2007; 28: 91–106.

87. Feige JN, Auwerx J. DisSIRTing on LXR and cholesterol metabolism. Cell Metab. 2007; 6: 343–345.

88. Peet DJ, Janowski BA, Mangelsdorf DJ. The LXRs: a new class of oxysterol receptors. Curr Opin Genet Dev. 1998; 8: 571–575.

89. Xu P, Li D, Tang X, et al. LXR agonists: new potential therapeutic drug for neurodegenerative diseases. Mol Neu- robiol. 2013; 48: 715–728.

90. Chang H-C, Guarente L. SIRT1 and other sirtuins in metabolism. Trends Endocrinol Metab. 2014; 25: 138–145.

91. Lee M-W, Chanda D, Yang J, et al. Regulation of hepatic gluconeogenesis by an ER-bound transcription factor, CREBH. Cell Metab. 2010; 11: 331–339.

92. Tang BL, Chua CEL. Is systemic activation of Sirt1 beneficial for ageing-associated metabolic disorders? Biochem Biophys Res Commun. 2010; 391: 6–10.

93. Puigserver P, Rhee J, Donovan J, et al. Insulin-regulated hepatic gluconeogenesis through FOXO1-PGC-1alpha interaction. Nature. 2003; 423: 550.

94. Zhou G, Myers R, Li Y, et al. Role of AMP-activated protein kinase in mechanism of metformin action. J Clin Invest. 2001; 108: 1167.

17 Advances in Biochemistry & Applications in Medicine 95. Bernier M, Paul RK, Martin-Montalvo A, et al. Negative regulation of STAT3 protein-mediated cellular respiration by SIRT1 protein. J Biol Chem 2011; 286: 19270–19279.

96. Li X, Kazgan N. Mammalian sirtuins and energy metabolism. Int J Biol Sci. 2011; 7: 575.

97. Zong H, Ren JM, Young LH, et al. AMP kinase is required for mitochondrial biogenesis in skeletal muscle in re- sponse to chronic energy deprivation. Proc Natl Acad Sci. 2002; 99: 15983–15987.

98. Fernandez-Marcos PJ, Auwerx J. Regulation of PGC-1α, a nodal regulator of mitochondrial biogenesis. Am J Clin Nutr. 2011; 93: 884S–890S.

99. Wu H, Kanatous SB, Thurmond FA, et al. Regulation of mitochondrial biogenesis in skeletal muscle by CaMK. Sci- ence. 2002; 296: 349–352.

100. Sun C, Zhang F, Ge X, et al. SIRT1 improves insulin sensitivity under insulin-resistant conditions by repressing PTP1B. Cell Metab. 2007; 6: 307–319.

101. Inoue H, Ogawa W, Asakawa A, et al. Role of hepatic STAT3 in brain-insulin action on hepatic glucose production. Cell Metab. 2006; 3: 267–275.

102. Bellet MM, Sassone-Corsi P. Mammalian circadian clock and metabolism–the epigenetic link. J Cell Sci. 2010; 123: 3837–3848.

103. Balsalobre A, Marcacci L, Schibler U. Multiple signaling pathways elicit circadian gene expression in cultured Rat-1 fibroblasts. Curr Biol. 2000; 10: 1291–1294.

104. Park I, Lee Y, Kim H-D, et al. Effect of Resveratrol, a SIRT1 Activator, on the Interactions of the CLOCK/BMAL1 Complex. Endocrinol Metab. 2014; 29: 379–387.

18